Tải bản đầy đủ (.pdf) (345 trang)

Meat Refrigeration-CRC Press (2002)_S.J. James, C.James

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (2.07 MB, 345 trang )


Meat refrigeration
S. J. James and C. James

Cambridge England


Published by Woodhead Publishing Limited, Abington Hall, Abington
Cambridge CB1 6AH, England
www.woodhead-publishing.com
Published in North America by CRC Press LLC, 2000 Corporate Blvd, NW
Boca Raton FL 33431, USA
First published 2002, Woodhead Publishing Ltd and CRC Press LLC
© The University of Bristol 2002
The authors have asserted their moral rights.
This book contains information obtained from authentic and highly regarded sources.
Reprinted material is quoted with permission, and sources are indicated. Reasonable efforts
have been made to publish reliable data and information, but the authors and the
publishers cannot assume responsibility for the validity of all materials. Neither the authors
nor the publishers, nor anyone else associated with this publication, shall be liable for any
loss, damage or liability directly or indirectly caused or alleged to be caused by this book.
Neither this book nor any part may be reproduced or transmitted in any form or by any
means, electronic or mechanical, including photocopying, microfilming and recording, or by
any information storage or retrieval system, without permission in writing from the
publishers.
The consent of Woodhead Publishing and CRC Press does not extend to copying for
general distribution, for promotion, for creating new works, or for resale. Specific permission
must be obtained in writing from Woodhead Publishing or CRC Press for such copying.
Trademark notice: Product or corporate names may be trademarks or registered
trademarks, and are used only for identification and explanation, without intent to infringe.
British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library.
Library of Congress Cataloging in Publication Data
A catalog record for this book is available from the Library of Congress.
Woodhead Publishing ISBN 1 85573 442 7
CRC Press ISBN 0-8493-1538-7
CRC Press order number: WP1538
Cover design by The ColourStudio
Typeset by SNP Best-set Typesetter Ltd., Hong Kong
Printed by TJ International, Padstow, Cornwall, England


Related titles from Woodhead’s food science, technology and
nutrition list:
Meat processing (ISBN 1 85573 583 0)
This major new collection summarises key developments in research,
from improving raw meat quality and safety issues to developments in
meat processing and specific aspects of meat product quality such as
colour, flavour and texture.
HACCP in the meat industry (ISBN 1 85573 448 6)
Following the crises involving BSE and E. coli the meat industry has
been left with an enormous consumer confidence problem. In order to
regain the trust of the general public the industry must establish and
adhere to strict hygiene and hazard control systems. HACCP is a
systematic approach to the identification, evaluation and control of food
safety hazards. It is being applied across the world, with countries such as
the USA, Australia, New Zealand and the UK leading the way. However,
effective implementation in the meat industry remains difficult and
controversial. This book is a survey of key principles and best practice,
providing an authoritative guide to making HACCP systems work
successfully in the meat industry.

Lawrie’s meat science (ISBN 1 85573 395 1)
This book remains a standard for both students and professionals in the
meat industry. It provides a systematic account of meat science from the
conception of the animal until human consumption, presenting the
fundamentals of meat science. This sixth edition incorporates the
significant advances in meat science which have taken place during the
past decade including our increasingly precise understanding of the
structure of the muscle, as well as the identification of the aberrations in
DNA which lead to the development of BSE syndrome in meat.
Details of these books and a complete list of Woodhead titles can be
obtained by:
᭿
᭿

visiting our web site at: www.woodhead-publishing.com
contacting Customer Services (e-mail: ; tel: +44 (0)1223 891358 ext. 30; fax: +44 (0)1223
893694; address: Woodhead Publishing Ltd, Abington Hall, Abington,
Cambridge CB1 6AH, England)


Contents

Part 1 Refrigeration and meat quality
1

Microbiology of refrigerated meat
1.1
Factors affecting the refrigerated shelf-life of meat
1.1.1 Initial microbial levels
1.1.2 Temperature

1.1.3 Relative humidity
1.2
Other considerations
1.2.1 Bone taint
1.2.2 Cold deboning
1.2.3 Hot deboning
1.3
Conclusions
1.4
References

3
4
4
6
11
12
13
14
15
16
16

2

Drip production in meat refrigeration
2.1
Biochemistry of meat
2.1.1 Structure of muscle
2.1.2 Changes after slaughter

2.1.3 Water relationships in meat
2.1.4 Ice formation in muscle tissues
2.2
Measurement of drip
2.3
Factors affecting the amount of drip
2.3.1 Animal factors
2.3.2 Refrigeration factors
2.3.3 Chilled storage

21
22
22
25
27
29
30
30
30
33
36


vi

Contents
2.4
2.5

Conclusions

References

40
41

3

Effect of refrigeration on texture of meat
3.1
Muscle shortening
3.1.1
Mechanism of shortening
3.1.2
Preventing shortening
3.2
Development of conditioning (ageing)
3.2.1
Mechanism of ageing
3.2.2
Prediction of tenderness
3.2.3
Consumer appreciation of ageing
3.2.4
Preslaughter factors
3.2.5
Pre-rigor factors
3.2.6
At chill temperatures
3.2.7
At frozen temperatures

3.2.8
At higher temperatures
3.3
Influence of chilling on texture
3.3.1
Lamb
3.3.2
Pork
3.3.3
Beef
3.4
Influence of freezing on texture
3.4.1
Lamb
3.4.2
Pork
3.4.3
Beef
3.5
Influence of thawing on texture
3.6
Conclusions
3.7
References

43
44
45
49
50

51
52
52
53
54
56
57
58
59
59
59
61
61
62
63
63
64
64
66

4

Colour changes in chilling, freezing and storage of meat
4.1
Meat colour
4.2
Factors affecting the colour of meat
4.2.1
Live animal
4.2.2

Chilling
4.2.3
Conditioning
4.2.4
Chilled storage
4.2.5
Freezing
4.2.6
Frozen storage
4.2.7
Thawing
4.2.8
Retail display
4.3
Conclusions
4.4
References

71
71
73
73
73
74
75
76
76
78
79
81

82

5

Influence of refrigeration on evaporative weight loss from
meat
5.1
Theoretical considerations

85
86


Contents
5.2

5.3
5.4
5.5

Weight loss in practice
5.2.1 Chilling
5.2.2 Chilled storage
5.2.3 Freezing and frozen storage
5.2.4 Retail display
Overall
Conclusions
References

vii

87
88
90
91
92
93
94
95

Part 2 The cold chain from carcass to consumer
6

Primary chilling of red meat
6.1
Introduction
6.2
Conventional chilling
6.2.1 Beef
6.2.2 Lamb, mutton and goat chilling
6.2.3 Pork
6.2.4 Chilling of offal
6.3
Novel systems with future potential
6.3.1 Accelerated chilling systems
6.3.2 Spray chilling
6.3.3 Immersion chilling
6.3.4 Ice bank chilling
6.3.5 Combined systems
6.3.6 Protective coatings
6.3.7 Hot boning

6.4
Conclusions
6.5
References

99
99
100
100
110
115
118
119
119
123
125
127
128
129
129
132
132

7

Freezing of meat
7.1
Freezing rate
7.2
Freezing systems

7.2.1 Air
7.3
Contact freezers
7.4
Cryogenic freezing
7.5
Freezing of specific products
7.5.1 Meat blocks
7.5.2 Beef quarters
7.5.3 Mutton carcasses
7.5.4 Offal
7.5.5 Small products
7.6
Tempering and crust freezing
7.6.1 Pork loin chopping
7.6.2 High speed ham slicing

137
137
140
140
142
144
145
145
145
146
146
147
149

149
150


viii

Contents

7.7
7.8
8

9

10

7.6.3
High speed bacon slicing
Conclusions
References

Thawing and tempering
8.1
Considerations
8.2
Quality and microbiological considerations
8.3
Thawing systems
8.3.1
Conduction

8.3.2
Electrical methods
8.3.3
Published thawing data for different meat cuts
8.3.4
Commercial practice
8.4
Tempering
8.4.1
Requirements for cutting and processing
equipment
8.4.2
Requirements for prebreaking
8.4.3
Microwave tempering
8.4.4
Commercial practice
8.5
Conclusions
8.6
References

150
155
155
159
160
161
163
166

166
168
176
178
178
179
182
185
186
187

Transportation
9.1
Sea transport
9.2
Air transport
9.3
Overland transport
9.3.1
Types of refrigeration system
9.3.2
Observations of transport
9.3.3
Problems particular to local delivery vehicles
9.3.4
Design and operation of local distribution
vehicles
9.4
Changes during transportation
9.5

Conclusions
9.6
References

191
191
193
193
194
195
197

Chilled and frozen storage
10.1 Storage life terms
10.2 Chilled storage
10.2.1 Unwrapped meat
10.2.2 Wrapped meat
10.2.3 Cooked products
10.3 Frozen storage
10.3.1 Oxidative rancidity
10.3.2 Prefreezing treatment
10.3.3 Freezing process
10.3.4 During frozen storage

207
207
208
209
211
214

216
216
218
220
221

198
202
204
204


Contents
10.4

Types of storage room
10.4.1 Bulk storage rooms
10.4.2 Controlled atmosphere storage rooms
10.4.3 Jacketed cold stores
Conclusions
References

224
224
225
225
225
226

Chilled and frozen retail display

11.1 Chilled display of wrapped meat and meat products
11.1.1 Factors affecting display life
11.1.2 Layout of chilled cabinet
11.1.3 Air curtain
11.1.4 Cabinet development
11.1.5 Computer modelling
11.1.6 Store conditions
11.2 Retail display of unwrapped meat and delicatessen
products
11.2.1 Types of cabinet
11.2.2 Appearance changes
11.2.3 Effects of environmental conditions
11.3 Retail display of frozen wrapped meat
11.3.1 Factors controlling display life
11.4 Overall cabinet design
11.4.1 Air circulation and temperatures
11.4.2 Effect of doors and lids
11.4.3 Effect of radiant heat
11.4.4 Measurement methods
11.5 Conclusions
11.6 References

231
231
232
233
234
235
236
236


10.5
10.6
11

12

ix

Consumer handling
12.1 Consumer attitudes to food poisoning
12.2 Shopping habits and transport from retail store to
the home
12.3 Refrigerated storage in the home
12.4 Temperatures in domestic food storage
12.5 Performance testing of domestic refrigerators
12.5.1 Performance of empty appliances
12.5.2 Performance of loaded appliances
12.5.3 Effect of loading with warm (20 °C) food
products
12.5.4 Effect of door openings
12.6 Performance testing of domestic freezers
12.7 Conclusions
12.8 References

237
238
238
239
241

241
244
245
246
247
247
248
249
251
252
252
255
256
262
263
263
264
264
265
267
269


x

Contents

Part 3 Process control
13


Thermophysical properties of meat
13.1 Chilling
13.1.1 Thermal conductivity
13.1.2 Specific heat
13.1.3 Enthalpies
13.2 Freezing, thawing and tempering
13.2.1 Ice content
13.2.2 Heat extraction
13.2.3 Thermal conductivity
13.2.4 Density
13.3 Mathematical models
13.4 Conclusions
13.5 References

273
274
274
274
276
277
277
277
278
280
280
280
281

14


Temperature measurement
14.1 Instrumentation
14.1.1 Hand-held digital thermometers
14.1.2 Temperature recorders
14.1.3 Time–temperature indicators
14.2 Calibration
14.3 Measuring temperature data
14.3.1 Contact non-destructive methods
14.3.2 Non-contact non-destructive methods
14.3.3 Contact destructive methods
14.3.4 Storage
14.3.5 Distribution
14.3.6 Retail
14.4 Interpreting temperature data
14.4.1 Example 1
14.4.2 Example 2
14.5 Conclusions
14.6 References

283
284
284
285
288
289
289
290
290
292
294

295
296
298
298
299
301
302

15

Specifying, designing and optimising refrigeration systems
15.1 Process specification
15.1.1 Throughput
15.1.2 Temperature requirements
15.1.3 Weight loss
15.1.4 Future use
15.1.5 Plant layout
15.2 Engineering specification
15.2.1 Environmental conditions
15.2.2 Room size

303
303
304
304
304
305
305
306
307

308


Contents
15.2.3 Refrigeration loads
15.2.4 Refrigeration plant capacity
15.2.5 Relative humidity
15.2.6 Ambient design conditions
15.2.7 Defrosts
15.2.8 Engineering design summary
Procurement
15.3.1 Plant design
Optimisation
15.4.1 Process definition
Conclusions

308
311
312
313
313
313
314
314
317
317
320

Secondary chilling of meat and meat products
16.1 Cooked meat

16.1.1 Legislation
16.1.2 Practical
16.1.3 Experimental studies
16.2 Pastry products
16.2.1 Commerical operations
16.2.2 Experimental studies
16.3 Solid/liquid mixtures
16.4 Process cooling
16.5 Cook–chill
16.5.1 Cook–chill guidelines
16.5.2 Practical cooling time data
16.5.3 Refrigeration problems in practice
16.6 Conclusions
16.7 References

321
322
322
323
324
328
328
329
330
332
332
333
334
336
338

338

Index

341

15.3
15.4
15.5
16

xi


Part 1
Refrigeration and meat quality


1
Microbiology of refrigerated meat

There are many pertinent texts on the microbiology of meats. The purpose
of this chapter is to examine briefly the types of micro-organisms and conditions that are of interest in relation to the refrigeration of meat and meat
products.
In a perfect world, meat would be completely free of pathogenic (food
poisoning) micro-organisms when produced. However, under normal
methods the production of pathogen-free meat cannot be guaranteed. The
internal musculature of a healthy animal is essentially sterile after slaughter
(Gill, 1979, 1980). However, all meat animals carry large numbers of different micro-organisms on the outer surfaces of the body and in the alimentary
tract. Only a few types of bacteria directly affect the safety and quality of

the finished carcass. Of particular concern are foodborne pathogens such
as Campylobacter spp., Clostridium perfringens, pathogenic serotypes of
Escherichia coli, Salmonella spp., and Yersinia enterocolitica.
In general, the presence of small numbers of pathogens is not a problem
because meat is normally cooked before consumption. Adequate cooking
will substantially reduce the numbers, if not completely eliminate all of the
pathogenic organisms present on the meat. Most meat-based food poisoning is associated with inadequate cooking or subsequent contamination after
cooking.The purpose of refrigeration is to reduce or eliminate the growth of
pathogens so that they do not reach levels that could cause problems.
Normally the growths of spoilage organisms limit the shelf-life of meat.
The spoilage bacteria of meats stored in air under chill conditions include
species of Pseudomonas, Brochothrix and Acinetobacter/Moraxella. In
general, there is little difference in the microbial spoilage of beef, lamb, pork
and other meat derived from mammals (Varnam and Sutherland, 1995).


4

Meat refrigeration

Meat is considered spoiled by bacteria when the products of their metabolic activities make the food offensive to the senses of the consumer (Gill,
1983). Therefore, the perception of a state of spoilage is essentially a subjective evaluation that will vary with consumer expectations. Few, however,
would not acknowledge that the appearance of slime, gross discoloration
and strong odours constitute spoilage.
‘Off’ odours are due to an accumulation of malodorous metabolic products, such as esters and thiols. Several estimations have been made of the
number of bacteria on meat at the point at which odour or slime becomes
evident and the mean is about 3 ¥ 107 cm-2 (Shaw, 1972).When active growth
occurs, the number of bacteria increases exponentially with time. Therefore,
a convenient measure of the growth rate is the time required for doubling
of numbers, often called the generation time. If this, for example, were one

hour, the number would increase two-fold in 1 h, four-fold in 2 h, eight-fold
in 3 h, and so on.
The bacterial safety and rate of spoilage depends upon the numbers and
types of micro-organisms initially present, the rate of growth of those microorganisms, the conditions of storage (temperature and gaseous atmosphere)
and characteristics (pH, water activity aw) of the meat. Of these factors, temperature is by far the most important.

1.1

Factors affecting the refrigerated shelf-life of meat

1.1.1

Initial microbial levels

1.1.1.1 Tissue sterility
For many years microbiologists believed that the tissues of healthy animals
normally contained bacteria (Reith, 1926; Ingram, 1972). These ‘intrinsic’
bacteria were the cause of phenomena such as ‘bone taint’. The cause of
bone taint is still questioned and will be discussed later. The prevailing view
of the majority of textbooks (Banwart, 1989; Varnam and Sutherland, 1995),
based in part on the work of Gill (Gill, 1979, 1980) is that the meat of a
healthy animal is essentially sterile. Low numbers of specific microorganisms, which have reached the tissues during the life of the animal, may
occur in the viscera and associated lymph nodes from time to time (Gill,
1979; Roberts and Mead, 1986). These are often pathogenic species, such as
Salmonella, and clostridia spores. The absence of bacteria appears to be due
to the continued functioning of the immune system in slaughtered animals.
Experiments with guinea pigs showed that the antibacterial defences of live
animals persisted for an hour or more after death and could inactivate
bacteria introduced during slaughter (Gill and Penney, 1979). Clearly, if
bacteria are thus inactivated there can be no multiplication, in deep tissue,

during carcass chilling irrespective of cooling rates.


Microbiology of refrigerated meat

5

1.1.1.2 Rigor mortis
The way in which animals are handled before slaughter will effect the biochemical processes that occur before and during rigor mortis. The resulting
metabolites influence the growth of micro-organisms on meat.
During the onset of rigor mortis, which may take up to 24 h, oxygen
stored in the muscle is depleted and the redox potential falls from above
+250 mV to -150 mV. Such a low redox value combined with the initial
muscle temperature of 38 °C provides ideal growth conditions for mesophilic micro-organisms. Stress and excitement caused to the animal before
slaughter will cause the redox potential to fall rapidly, possibly allowing
proliferation of such micro-organisms before cooling (Dainty, 1971).
Concurrent with the fall in redox potential is a fall in pH from an initial
value in life of around 7 to a stable value around 5.5, the ‘ultimate pH’. This
is due to the breakdown of glycogen, a polysaccharide, to lactic acid in the
muscle tissue. Lactic acid cannot be removed by the circulation nor oxidised, so it accumulates and the pH falls until the glycogen is all used or
the breakdown stops. The pH has an important role in the growth of microorganisms, the nearer the pH is to the ultimate value, the more growth is
inhibited (Dainty, 1971). Stress or exercise before slaughter can deplete an
animal’s glycogen reserves, consequently producing meat with less lactic
acid and a relatively high ultimate pH, this gives the meat a dark, firm, dry
(DFD) appearance. Alternative terms are ‘dark cutting’ and ‘high-pH
meat’. The condition occurs in pork, beef and mutton, but is of little economic importance in the latter (Newton and Gill, 1981). DFD meat provides conditions that are more favourable for microbial growth than in
normal meat. The microbiology of DFD meat has been comprehensively
reviewed by Newton and Gill (1981).
Glucose is the preferred substrate for growth of pseudomonads, the
dominant bacteria in meat stored in air at refrigerated temperatures. Only

when glucose is exhausted do they break down amino acids, producing the
ammonia and sulphur compounds that are detectable as spoilage odours
and flavours. In meat containing no glucose, as is the case with some DFD
meat, amino acids are broken down immediately and spoilage becomes
evident at cell densities of 6 log10 cfu cm-2 (colony forming units per centimetre squared). This is lower than in normal meat, where spoilage
becomes apparent when numbers reach ca. 8 log10 cfu cm-2. Thus, given the
same storage conditions, DFD meat spoils more rapidly than normal-pH
meat. There is no evidence that the spoilage of pale, soft, exuding (PSE)
meat is any different to that of normal meat (Gill, 1982). There is little significant difference in pH or chemical composition between PSE and normal
meat.
1.1.1.3 Surface contamination
Initial numbers of spoilage bacteria on carcasses significantly affect shelflife. With higher numbers, fewer doublings are required to reach a spoilage


6

Meat refrigeration

level of ca. 108 organisms cm-2. For example, starting with one organism
cm-2, 27 doublings would be needed; for 103 organisms cm-2 initially, the
number of doublings is reduced to 17.
Contamination of carcasses may occur at virtually every stage of slaughtering and processing, particularly during flaying and evisceration of redmeat animals and scalding, and mainly affects the surface of the carcass.
Sources of contamination have been reviewed by James et al. (1999).
Hygienic handling practices should ensure that total viable counts on the
finished carcass are consistently 103–104 organisms cm-2 or lower for red
meats. Bad practices can cause counts to exceed 106 organisms cm-2.
With red meats, carcasses of good microbial quality are obtained by
1
2
3


preventing contamination from the hide;
avoiding gut breakage;
the adoption of good production practices that include more humane
practices throughout the slaughtering system.

The effectiveness of chemical and physical decontamination systems for
meat carcasses has been reviewed by James and James, (1997) and James
et al. (1997). Commercial systems using steam have been introduced into
the USA and are claimed to reduce the number of bacteria on the surface
of beef carcasses to below 1 log10 cfu cm-2 (Phebus et al., 1997).

1.1.2 Temperature
Micro-organisms are broadly classified into three arbitrary groups (psychrophiles, mesophiles and thermophiles) according to the range of temperatures within which they may grow. Each group is characterised by three
values: the minimum, optimum and maximum temperatures of growth.
Reduction in temperature below the optimum causes an increase in generation time, i.e. the time required for a doubling in number. It is an accepted
crude approximation that bacterial growth rates can be expected to double
with every 10 °C rise in temperature (Gill, 1986). Below 10 °C, however, this
effect is more pronounced and chilled storage life is halved for each 2–3 °C
rise in temperature. Thus the generation time for a pseudomonad (a
common form of spoilage bacteria) might be 1 h at 20 °C, 2.5 h at 10 °C, 5 h
at 5 °C, 8 h at 2 °C or 11 h at 0 °C. In the usual temperature range for chilled
meat, -1.5–+5 °C, there can be as much as an eight-fold increase in growth
rate between the lower and upper temperature. Storage of chilled meat at
-1.5 ± 0.5 °C would attain the maximum storage life without any surface
freezing.
Meat stored above its freezing point, ca. -2 °C, will inevitably be spoiled
by bacteria. Obviously, the nearer the storage temperature of meat
approaches the optimum for bacterial growth (20–40 °C for most bacteria)
the more rapidly the meat will spoil.Work of Ayres (1960) compared the rate

of increase in bacterial number on sliced beef stored at 0, 5, 10, 15, 20 and


Microbiology of refrigerated meat

7

25 °C. The meat developed an off odour by the third day at 20 °C, the tenth
day at 5 °C and the 20th day at 0 °C. Similar data has been reported by other
workers.They clearly demonstrate the effectiveness of refrigeration in reducing the rate of increase in bacterial numbers and extending shelf-life.
As bacteria generally grow more rapidly than fungi, mould spoilage of
meat is thought to develop only when competing bacteria are inhibited.
Temperature is usually assumed to be the critical factor, mould spoilage
being typically associated with frozen meat. It has been generally accepted
that moulds can develop on meat at temperatures as low as -10 or -12 °C.
There is some evidence that this is an exaggeration and that for practical
purposes the minimum temperature for mould growth on meat should be
taken to be ca. -5 °C (Lowry and Gill, 1984). It is further thought that
surface desiccation, rather that temperature, is the factor that inhibits bacterial growth. If this is the case then mould growth on frozen meats is indicative of particularly poor temperature control.
Many factors influence the growth and survival of micro-organisms in
meat during freezing and frozen storage. However, the main factor affecting the growth of micro-organisms during freezing is the availability of
water. Until the temperature is reduced below the minimum temperature
for growth, some micro-organisms have the potential to multiply. While
most of the water in meat is turned to ice during freezing, there is always
some free liquid water available, 26% at -5 °C, 18% at -10 °C, 14% at
-18 °C, 10% at -40 °C (Rosset, 1982). The transformation of water into ice
significantly modifies the growth environment for micro-organisms, since
solutes become concentrated in the remaining free water to the level that
microbial growth is inhibited. Below the freezing point of the meat, the
water activity is progressively reduced preventing microbial growth (Fig.


Water activity (aw)

1.0

0.9

0.8

0.7
–30

–20

–10

0

Temperature (°C)

Fig. 1.1 Water activities (aw) of meat at various sub-freezing temperatures (source:
Leistner and Rödel, 1976).


8

Meat refrigeration

Table 1.1 Minimum and optimum growth temperatures for pathogens associated
with red meats


Campylobacter spp.
Clostridia perfringens
Pathogenic Escherichia
coli strains
Salmonella spp.
Listeria monocytogenes
Yersinia enterocolitica

Minimum temperature
(°C)

Optimum temperature
(°C)

30
12
7

42–43
43–47
35–40

5
0
-2

35–43
30–37
28–29


Source: Mead and Hinton, 1996.

1.1). The greatest reduction in the microbial load occurs during, or shortly
after, freezing itself. During frozen storage, the numbers are gradually
reduced further.
1.1.2.1 Pathogenic organisms
A number of bacterial pathogens capable of causing food poisoning in
humans are known to contaminate red meat. Those of most importance
are Campylobacter spp., Clostridium perfringens, pathogenic serotypes of
Escherichia coli (principally E. coli O157:H7), Salmonella spp. and Yersinia
enterocolitica (Nottingham, 1982; Anon, 1993; Mead and Hinton, 1996). Listeria monocytogenes is commonly associated with meat, but its public health
significance in relation to raw meat is unclear (Mead and Hinton, 1996).
The essential characteristics of pathogenic micro-organisms can be found
in numerous texts.
Minimum and optimum growth temperatures for pathogens commonly
associated with red meat are show in Table 1.1. Some pathogens, such as
L. monocytogenes, are capable of growth at chill temperatures below 5 °C.
These are often cited as being of particular concern in relation to refrigerated meats since refrigeration can not be relied on to prevent growth
(Doyle, 1987). On the other hand, psychrotrophic pathogens are not particularly heat resistant and adequate cooking should be sufficient to destroy
any such pathogens. Illnesses caused by L. monocytogenes and E. coli are
often due to inadequate cooking before ingestion.
1.1.2.2 Spoilage organisms
The number of types of micro-organisms capable of causing food spoilage
is very large and it is not possible to discuss them in any detail in this text.
Depending on the initial microflora and the growth environment, only a
few species of the genera Pseudomonas, Acinetobacter, Moraxella,
Lactobacillus, Brochothrix and Alteromonas, and of the family



Microbiology of refrigerated meat

9

Enterobacteriaceae are significantly represented in most spoilage
microflora of chilled meats (Bell and Gill, 1986).
The micro-organisms that usually spoil meat are psychrotrophs, i.e. they
are capable of growth close to 0 °C. Only a small proportion of the initial
microflora on meat will be psychrotrophs; the majority of micro-organisms
present are incapable of growth at low temperatures. As storage temperature rises the number of species capable of growth will increase.
1.1.2.2.1 Spoilage of chilled meat
The spoilage of chilled meat stored in air is dominated by Gram-negative,
psychrotrophic, aerobic rod-shaped bacteria. Although a wide range of
genera are present on meat, only Pseudomonas, Acinetobacter and
Psychrobacter species are normally of importance (Dainty and Mackey,
1992). Of these, species of Pseudomonas are of greatest importance (Gill,
1986). Pseudomonas spp. typically account for >50% of the flora and sometimes up to 90% (Dainty and Mackey, 1992).
Other bacteria are present in small numbers and may occasionally form
a significant part of the microflora. Brochothrix thermosphacta appears to
be of more importance on pork and lamb than on beef especially on fat
where the pH value is generally higher, and at temperatures above 5 °C
(Gill, 1983; Varnam and Sutherland, 1995).
Species of both Micrococcus and Staphylococcus are present on meat
stored in air but their significance is generally considered limited under
refrigerated storage. Psychrotrophic members of the Enterobacteriaceae,
including Serratia liquefaciens, Enterobacter agglomerans and Hafnia alvei
are also common at low levels (Dainty and Mackey, 1992). These organisms
become of greater importance at temperatures of 6–10 °C, but Pseudomonas spp. usually remain dominant (Varnam and Sutherland, 1995).
Yeasts and moulds are considered by many to be of limited importance
in modern practice (Varnam and Sutherland, 1995). Moulds were of historic

importance on carcass meat stored for extended periods at temperatures
just above freezing.
1.1.2.2.2 Spoilage of chilled packaged meat
Large vacuum packs usually contain ca. 1% O2 that in theory will support
the growth of pseudomonads (Varnam and Sutherland, 1995). Continuing
respiration, however, by the meat rapidly depletes oxygen (O2) and
increases the carbon dioxide (CO2) concentration to ca. 20%. Pseudomonas
spp. are usually unable to grow under such conditions. In general conditions
vacuum packs favour lactic acid bacteria (LAB), although there may also
be significant growth of Br. thermosphacta, ‘Shewanella putrefaciens’ (formally Altermonas putrefaciens) and the Enterobacteriaceae. Under anaerobic conditions LAB have a considerable advantage in growth rate over
competing species of facultative anaerobes (Fig. 1.2). The predominant
LAB are homofermentative species of Lactobacillus, Carnobacterium and


10

Meat refrigeration
Br. thermosphacta
Enterobacter
Lactobacillus

Generation time (h)

60
50
40
30
20
10
0

15

10

5

2

Temperature (°C)

Fig. 1.2 Effect of temperature on the rates of anaerobic growth of bacteria on meat
slices (source: Newton and Gill, 1978).

Leuconostoc. Lactococcus spp. are much less common. LAB are able to
grow at low temperatures and low O2 tensions and tolerate CO2. Psychrophilic species of Clostridium have been recognised as a significant
potential problem (Varnam and Sutherland, 1995).
Both Br. thermosphacta and Sh. putrefaciens are favoured by high pH
values. Sh. putrefaciens is unable to grow below pH 6.0 during storage at
low temperatures, whereas Br. thermosphacta is unable to grow anaerobically below pH 5.8 (Gill, 1983). At temperatures below 5 °C, Enterobacteriaceae are inhibited in vacuum packs by CO2, low pH and lactic acid. At
higher temperatures and pH values, CO2 is markedly less inhibitory and
growth is possible, in particular by Serratia liquefaciens and Providencia spp.
(Varnam and Sutherland, 1995).
The predominant type of spoilage in vacuum-packed chilled meat is
souring (Sofos, 1994; Varnam and Sutherland, 1995). This is not normally
detectable until bacterial numbers reach 8 log10 cfu cm-2 or greater. The
exact cause of such spoilage is unknown, but is assumed to result from lactic
acid and other end-products of fermentation by dominant LAB. High-pH
(DFD) vacuum-packed meat spoils rapidly and involves the production of
large quantities of hydrogen sulphide (H2S) by Sh. putrefaciens and Enterobacteriaceae (Gill, 1982; Varnam and Sutherland, 1995). Characteristic
‘greening’ occurs owing to H2S combining with the muscle pigment to

give green sulphmyoglobin; the meat also develops putrid spoilage odours
(Gill, 1982).
Packaging in various gaseous atmospheres has been used as an alternative to vacuum packing. The intention has been to preserve the fresh meat
colour and to prevent anaerobic spoilage by using high concentrations of


Microbiology of refrigerated meat

11

oxygen (50–100%) along with 15–50% carbon dioxide to restrict the growth
of Pseudomonas and related species (Nottingham, 1982). The microflora of
meat stored in commercially used modified atmosphere packs (MAP) is in
general similar to that of vacuum packs (Varnam and Sutherland, 1995). At
temperatures below 2 °C, LAB are dominant, Leuconostoc spp. being the
most important. Br. thermosphacta, Pseudomonas spp. and Enterobacteriaceae are more prevalent in MAP (modified atmosphere packs) than
vacuum packs at storage temperatures ca. 5 °C, rather than 1 °C. Br. thermosphacta is relatively CO2 tolerant and the presence of O2 permits growth
of this bacterium at pH values below 5.8. Prior conditioning in air favours
the growth of these bacteria, they are also more prevalent in pork than
other meats (Dainty and Mackey, 1992). The spoilage of meat in MAP may
involve souring similar to that in vacuum-packed meat. Other characteristics include ‘rancid’ and ‘cheesy’ odours. Chemical rancidity does not appear
to be primarily involved and souring is probably caused by the metabolites
of LAB or Br. thermosphacta (Varnam and Sutherland, 1995).
1.1.2.2.3 Spoilage of frozen meat
Micro-organisms do not grow below ca. -10 °C, thus spoilage is only normally relevant to handling before freezing or during thawing. In these
contexts, frozen meats behave like their unfrozen counterparts, although
growth rates may be faster after thawing, owing to drip.
Although Salmonella, Staphylococci, and other potential pathogens can
survive freezing and frozen storage, the saprophytic flora (spoilage bacteria) tend to inhibit their growth (Varnam and Sutherland, 1995). During
freezing and thawing of food, the temperature favours the growth of

psychrophilic organisms, most of which are spoilage organisms. Hence,
in nearly all cases, if a frozen product is mishandled, spoilage is apparent
before the food becomes a health hazard.
In the past, carcass meats were imported at temperatures of -5 to
-10 °C. At these temperatures there were problems with the growth of psychrotrophic moulds such as strains of Cladosporium, Geotrichum, Mucor,
Penicillium, Rhizopus and Thamnidium, causing ‘whiskers’ or ‘spots’ of
various colours depending on the species. Since little meat is now stored
at these temperatures mould spoilage is largely of historic importance.
Despite this, many meat microbiology textbooks continue to discuss this
subject in great detail.

1.1.3 Relative humidity
Historically low relative humidities (RH) have been recommended to
extend shelf-life. Schmid (1931) recommended a storage temperature for
meat of 0 °C and an RH of 90%. Haines and Smith (1933) later demonstrated that lowering the RH is more effective in controlling bacterial
growth on fatty or connective tissue than on lean meat. This was due to a
slower rate of diffusion of water to the surface. Low RH was more


12

Meat refrigeration

effective, therefore, in reducing microbial spoilage of carcass meat than of
small lean joints.
Micro-organisms normally grow in foods in the equivalent of a nutrient
solution and the availability of water in this solution is one of the factors
normally controlling growth. The term used for physiologically available
water is ‘water activity’ (aw). By definition
aw = P P0

Where P and P0 are the vapour pressures of the solution and of the pure
solvent, respectively. The above equation also defines the relative humidity
(RH) of the vapour phase in equilibrium with the solution. The RH (%) of
an atmosphere in equilibrium with a food would be 100 times the aw of the
food.
If the RH of the atmosphere corresponds to an aw lower than that of
meat then the meat surface will lose water to the atmosphere and the aw
will fall. In practice, the aw of the surface of meat will not fall to the corresponding RH value of the atmosphere, because water lost from the surface
is partially replaced by diffusion of water from the interior. However, the
basis of the efficacy of low relative humidities in extending shelf-life is the
reduction of the aw of the meat surface to a level inhibiting the growth of
psychrophilic bacteria.
The aw of lean meat is of the order of 0.993 (Scott, 1936) and offers ideal
growth conditions for micro-organisms. Scott (1936) and Scott and Vickery
(1939) established that the important meat spoilage bacteria are unable to
grow on meat at temperatures below 4 °C if the aw is less than 0.96. This
occurs when the water content has fallen from about 300 to about 85 g
of water/100 g of dry matter. They recommended that in chilled storage
the maximum values of the mean RH at air speeds of 0.15, 0.45, 0.70 and
0.90 m s-1 should be approximately 72, 85, 88 and 90%, respectively. These
conditions would maintain the water content at the surface of the meat
at or near inhibitory levels for bacteria. If conditions during the cooling
stage were unsatisfactory, they recommended that for any given air speed
the relative humidity would need to be lower to prevent bacterial growth.
These recommendations were made for long-term storage during shipment from Australia to the UK. Higher relative humidities may be used if
only short-term storage is the aim. This work remains the basis for the usual
recommendation to operate meat chillers between 85–95%. The actual RH
used will depend, of course, upon the air speed, the type of meat, the length
of storage required and the temperature of storage.


1.2

Other considerations

Legislation and recommendations for cooling of meat are believed to be
based on clear microbiological criteria. However, there are a lack of data


Microbiology of refrigerated meat

13

to support recommendations on the avoidance of bone taint, and on chilling rates for hot and cold boning.

1.2.1 Bone taint
It has been said (Moran and Smith, 1929) that ‘possibly the strongest argument for the rapid cooling of beef immediately after death is that it reduces
the possibility of bone taint’. Twenty years earlier in the USA, Richardson
and Scherubel (1909) had concluded that, to control the condition, the
carcass should be cooled to 4 °C or below at the centre of the round
(hindquarter) within 48 h of slaughter.
Bone taint has long been regarded as evidence for the presence of intrinsic bacteria. This view is diminishing, but the exact nature of what constitutes bone taint remains undefined (Nottingham, 1982; Shaw et al., 1986).
A wide range of bacteria has been implicated in the past. Varnam and
Sutherland (1995) in their book on meat report that halophilic Vibrio
species and Providencia are now most commonly attributed to this condition, while work by De Lacy et al. (1998) demonstrated that some strains
of psychrotrophic Clostridium spp. have the potential to cause bone taint.
Bone taint in beef is usually localised in the region of the hip joint and
is manifested by a ‘typical sewage type odour’ or ‘putrefactive sulphide-type
odours’. This is referred to as ‘souring’ in the American literature, which is
only detected when the hindquarter is divided (Thornton, 1951). Taints
in pork products appear to differ from bone taint in beef, occurring most

often after processing into cured hams or gammons (Jensen and Hess, 1941;
Haines, 1941).
The aetiology of bone taint remains obscure.The bacteria associated with
bone taint are supposed to have their origin in the bloodstream at death
and the infection starts in the blood vessels of the marrow of the femur
(Callow and Ingram, 1955). How they get into the blood supply is not established. One possibility is that unusually large numbers of bacteria are
introduced at slaughter, for example, by the use of slaughter instruments
contaminated with faeces (Jensen and Hess, 1941; Mackey and Derrick,
1979). Such massive contamination might be sufficient to overwhelm the
immune system leading to survival of a few cells (Gill and Penney, 1979).
Alternatively, bacteria may originate from undetected infection, for
example of joints. The lymph nodes have also been implicated as centres of
infection (Lepovetsky et al., 1953; Cosnett et al., 1956; Nottingham, 1960).
Surprisingly, in view of the early recognition of the importance of
temperature, there are apparently no definitive data on the cooling rate
required to assure freedom from bone taint in the various species of meat
animal. In all examples studied in cattle and sheep (i.e. ruminants) it is
agreed that bone taint will only manifest itself if cooling of the carcass postmortem is insufficiently rapid (Kitchell, 1972). Some conditions under
which bone taint has occurred are given in Table 1.2. With pigs, on the other


14

Meat refrigeration

Table 1.2

Cooling conditions under which bone-taint in beef has been reported

Reference

Moran and Smith (1929)
Haines and Scott (1940)
De Lacy et al. (1998)

Deep muscle temperature
(°C)

Time post-mortem
(h)

22
17
>18
20

24
48
40 or longer
20

hand, Jensen and Hess (1941) reported that, even under optimum cooling
conditions, 6–7.5% of several thousand hams exhibited various forms of
‘souring’. Unfortunately, the cooling data furnished in support of this statement include no temperatures measured at the hip joint.
From time to time bone taint of carcasses is still reported within the
industry. The rarity of bone taint may reflect a requirement for coincident
circumstances, each of which is uncommon. Microbial survival and growth
would be favoured by very high levels of contamination, possible weakening of the antimicrobial defences (as can occur in haemorrhagic shock), high
muscle pH and slow cooling. To allow for the infrequent occurrence of
bacteria in deep tissues it is thus prudent to cool carcasses promptly after
slaughter. The temperature of the deep leg should be brought below 15 °C

within 24 h.

1.2.2 Cold deboning
Council Directives of the EU (Council Directive No. 64/433/ECC, 1964;
Council Directive No. 83/90/ECC, 1983) stipulate that carcass meat must be
chilled immediately after post-mortem inspection. Chilling must continue
to an internal temperature of 7 °C before cutting or transportation can take
place. This requirement, aimed at preventing growth of salmonellas, has
caused problems in the meat industry. To allow deboning 24 h post-mortem,
the outer portions of the carcass may have to be cooled at a rate causing
toughening due to ‘cold shortening’ and surface fat may become very hard
and difficult to handle. The 1964 Council Directive also stipulated a
maximum cutting room temperature of 10 °C, which was increased to 12 °C
in the 1983 amendments to that Directive.
To judge the need for such a stringent regulation the effect of temperature on the growth of salmonellas on meat has been defined. Mackey et al.
(1980) used observed generation times of salmonellas on beef surfaces
maintained at high RH values to calculate the maximum extent of growth
during storage for different times at temperatures between 10 and 15 °C.
Smith (1985) produced tables of lag and generation times which can be used
to determine the length of time raw chilled meat can be held at temperatures between 10 and 40 °C without an increase in salmonella numbers.


Microbiology of refrigerated meat

15

Because of their studies, Mackey et al. (1980) concluded that under practical conditions of cutting and packaging that takes only 2–3 h, a meat
temperature of 10 °C would be entirely adequate to ensure no significant
multiplication of salmonella. Smith (1985) concluded that cutting rooms
could be maintained above 10 °C provided carcasses are processed

promptly and meat is not allowed to accumulate in the cutting room.

1.2.3 Hot deboning
Hot deboning has received much attention in recent years because of its
potential to streamline butchery, packaging and chilling. As the name
implies the carcass is deboned while hot and the meat is chilled in vacuum
packs or cartons. In principle, this could increase the risk of microbial
growth because there is no surface drying, and some contaminated surfaces
are deep in the meat where they will cool more slowly.
The possible microbiological effects of hot deboning have been investigated either by monitoring growth following natural contamination or by
prediction of growth based on data obtained following inoculation. Natural
contamination experiments have shown that it is possible to produce hot
deboned meat of the same microbiological quality as cold deboned meat
in terms of total counts and numbers of mesophilic pathogenic bacteria
(Taylor et al., 1980). Based on natural contamination experiments, Fung
et al. (1981) recommended chilling to 21 °C within 3–9 h after packaging
with continuous chilling to below 10 °C within 24 h. The validity of using
natural contamination to monitor the growth of pathogens on hot deboned
meat has, however, been questioned (Grau, 1983) because initial numbers
are very low, often undetectable, and heterogeneous in distribution. As an
alternative, prediction of growth at different cooling rates can be calculated
from an equation derived from observations on the growth of the organism on meat following inoculation (Herbert and Smith, 1980). Cooling rates
recommended on this basis are more rapid (e.g. cool to 8 °C from an initial
30 °C within 6 h of the commencement of boning) than those based on
natural contamination experiments.
Microbiologists have yet to agree whether the observational (natural
contamination) or predictive approach is more appropriate in defining the
effect of different cooling rates on the extent of microbial growth. Prediction using mathematical models based on data from inoculation experiments avoids the need to perform time-consuming tests on each cooling
rate in question. In some instances, where factors affecting growth are fully
understood, it may even be possible to predict growth with reasonable precision from growth rate data obtained in laboratory media, as demonstrated

by Gill (1984) for E. coli in tub-packed livers. The accuracy of all predictive
models must, however, be confirmed as far as is possible by comparison with
observed growth at a selection of cooling rates following natural contamination, as performed by Gill (1984).


×